Sie sind auf Seite 1von 6

SSC 5073

PERGAMON Solid State Communications 113 (2000) 553–558


www.elsevier.com/locate/ssc

In situ Raman scattering studies of the amorphous and crystalline


Si nanoparticles
A.A. Sirenko a,*, J.R. Fox a, I.A. Akimov a,b, X.X. Xi a, S. Ruvimov c, Z. Liliental-Weber c
a
Department of Physics, The Pennsylvania State University, University Park, PS 16802, USA
b
A.F. Ioffe Physico-Technical Institute, Russian Academy of Sciences, St Petersburg 194021, Russian Federation
c
Lawrence Berkeley Laboratory, Berkeley, CA 94720, USA
Received 10 September 1999; received in revised form 26 October 1999; accepted 24 November 1999 by J. Kuhl

Abstract
We report on in situ studies of the vibrational properties of Si nanoparticles and ultrathin layers grown by dc magnetron
sputtering in ultrahigh vacuum on amorphous MgO and Ag buffer layers. The average thickness of the Si layers ranged from
monolayer coverage up to 200 Å. Transmission electron microscopy has been used to determine size and shape of the Si
nanoparticles. Changes in the phonon spectra of Si nanoparticles during the crystallization process have been studied by
interference enhanced Raman scattering technique. Marked size-dependences in the phonon density of states and the relaxation
of the k-vector conservation with decrease in size of the Si nanoparticles have been detected. The transition between crystalline-
and amorphous-like behavior takes place in the particles with an average number of Si atoms equal to …7 ^ 2† × 102 : q 2000
Elsevier Science Ltd. All rights reserved.
Keywords: A. Nanostructures; D. Phonons; E. Inelastic light scattering

1. Introduction due to a difference in the bulk dispersions and the dielectric


properties of the nanoparticles and the surrounding media.
As the size of condensed matter systems is reduced in one This effect has been experimentally observed in a number of
or more dimensions to nanoscale levels, the electron and semiconductor nanostructures by means of Raman scatter-
phonon states are influenced due to confinement. The ing [5,16,17,18,19,20]. However, previous Raman scatter-
discovery of visible luminescence at room temperature ing studies were mostly focused on samples with Si particles
from porous Si [1,2] increased interest in the modification within a matrix or with ligands, such as polycrystalline
of electronic and vibrational properties of silicon nano- hydrogenated Si [5], Si-rich SiO2 films [6], prepared by
crystals. The prospect of realizing light-emitting devices pulverizing bulk Si [7], porous Si [17–19,21], or Si prepared
stimulated intensive investigations of various nanometer- by thermal evaporation [22]. In these systems, the influence
sized Si-based structures using different kinds of preparation of a matrix or surface contamination on the phonon spectra
methods [2–8]. Recent experimental studies of nanocrystal- cannot be neglected. Thus, in the analysis of the observed
line Si (nc-Si) demonstrated substantial changes in the shifts and broadening of the optical phonon lines, it was
luminescence properties and electronic bands, which were difficult to distinguish between effects due to phonon
attributed to quantum confinement caused by the restricted confinement, modification of the surface states, and
size of the particles and a breakdown of k-vector conserva- matrix-induced stress. While all the above mentioned effects
tion [9–12]. are possible, the phonon states of isolated Si nanoparticles
Phonon states are also modified in semiconductor nanos- have been relatively unexplored experimentally and publi-
tructures (or quantum dots) from those of the bulk. Theory cations on in situ measurements of nc-Si are rare [23].
predicts confinement of the optical and acoustic vibrations In this paper, we report on in situ Raman studies of ultra-
[13–15], which can be described macroscopically as being small Si particles and thin films grown in ultrahigh vacuum
(UHV). The phonon states of the Si nanoparticles free of
* Corresponding author. chemisorbed species as well as changes induced by the
0038-1098/00/$ - see front matter q 2000 Elsevier Science Ltd. All rights reserved.
PII: S0038-109 8(99)00539-6
554 A.A. Sirenko et al. / Solid State Communications 113 (2000) 553–558

determined that the Si clusters contain about …70 ^ 20† × d


(Å) atoms for d less than 50 Å. For the samples with d
more than 50 Å, no nanoparticle formation has been
observed and their properties are the same as of thin
films. It is similar to the growth regime of Ge nanoparticles
with a hemispherical shape investigated previously with the
same apparatus [24].
The Raman measurements were performed in situ at room
Fig. 1. TEM cross-section image of the recrystallized 30 Å sample
grown on Ag buffer. temperature and at a pressure of 2 × 10210 Torr in the UHV
system. The main experimental problem in the system of
two-dimensional composition of Si nanoparticles was a
crystallization process and hydrogen termination will be weak Raman signal due to extremely small scattering
discussed and vibration properties of the Si nanoparticles volume. For Si nanoparticles grown on MgO layers, we
grown on dielectric and metallic substrates will be improved the situation by using interference enhanced
compared. Results of in situ electron energy loss spectro- Raman scattering (IERS) [25,26]. The Raman signal from
scopy (EELS) will be used for the characterization of the the Si layer was increased by a factor of 10 in comparison
optical gap of Si nanoparticles. with similar samples grown on Ag due to constructive inter-
ference in a bilayer structure. We found that in combination
with multichannel detection [27], IERS can be sensitive
2. Experiment even to a submonolayer coverage. Recently this technique
provided useful information about the vibrations of surface
Ultrathin layers of Si were deposited on amorphous MgO and near-surface atoms of isolated Ge nanoparticles [24] and
or Ag buffers at room temperature by dc magnetron sputter- carbon nanocrystalline clusters and films [28]. In samples
ing in a UHV chamber. A power of 6 W/cm 2 was applied to grown on MgO buffer-layers, the interference enhancement
the Si sputter target. The background pressure of Ar was of the laser electric field close to the surface (the position of
kept around 4 mTorr during the Si growth process. Before the Si layer) was achieved by utilizing a bilayer structure,
deposition of Si, the substrates were annealed at 5008C. The which consists of a transparent dielectric film (MgO) depos-
growth rate of Si, calibrated by both ex situ ellipsometry and ited on a metallic layer (Ag) [25,26]. The entire structure
profilometry, was about 0.5 Å/s. The average thickness of was grown on a Si substrate. The thickness of the amorphous
the Si layers, d, ranged from monolayer coverage up to Ag, which works in a bilayer structure as a light reflector,
200 Å. In the following, the Si samples will be identified was about 4000 Å. This is thick enough to screen comple-
by their value of d. To crystallize Si ultrathin layers, the tely any possible Raman signal from the Si substrate. The
samples were heated up to about 5008C for 30 min in a MgO thickness of 500 Å was calculated to give optimal
UHV. The temperature of the sample was estimated by interference enhancement at the exciting wavelength. The
means of both a thermocouple attached to the sample holder choice for the dielectric buffer layer (MgO) was determined
and a pyrometer. by the two following reasons. First, the luminescence and
The grown ultrathin Si layers consist of nanometer-size Raman signal of MgO is weak at the green light excitation
clusters of a hemispherical shape formed on the substrate. and, second, the analysis of the XPS spectra is easier than,
Their average size and concentration varied with d. Ex situ e.g. in the case of a SiO2 buffer, which also contains Si
transmission electron microscopy (TEM) was employed to atoms.
characterize the shape and size of Si clusters in the samples Raman spectra were measured with a resolution of
grown on Ag buffer layers. Ag films with Si clusters were 0.8 cm 21 using a SPEX Triplemate spectrometer equipped
mechanically removed from the Si substrate, first, and then with a charge-coupled device (CCD) detector cooled with
folded so that the Si clusters on the edge can be studied by liquid nitrogen. The 5145 Å line (2.4 eV) of an Ar 1-ion
TEM. Two Si clusters are seen in a cross-section image of laser was used for excitation. The pumping power
the 30 Å sample annealed at 5008C (Fig. 1). The Si clusters density was about 1 W/cm 2. Both the intensity and
in this sample have a 70–100 Å base width and are 30–40 Å linear polarization of the Raman signal were analyzed
high. The Si layer with a thickness of about 10, covers Ag in the conventional backscattering configuration (laser
buffer between quantum dots. These data are in a quantita- beam and scattered light are perpendicular to the sample
tive agreement with the calibrated average amount of depos- plane). In the following, symbols VV and VH denote the
ited Si. Our results with in situ core-level X-ray scattering configurations with the polarization of the Raman
photoemission (XPS) show that at d # 50 A;  Si forms signal being parallel and perpendicular to the polarization of
nanoscale islands on MgO substrates as well. For thicker the exciting beam, respectively. EELS spectra were
layers of Si, the characteristic MgO signal in the XPS spec- measured with incident electron energy of 50 eV. An
tra quickly disappears, corresponding to a complete cover- LK2000 spectrometer was operated with 5 eV pass energy
age of the buffer-layer surface with Si atoms. We yielding a resolution of approximately 0.03 eV.
A.A. Sirenko et al. / Solid State Communications 113 (2000) 553–558 555

Fig. 3. The HWHM d TO and position D TO of the TO Raman peak as


functions of the average thickness of the amorphous Si layers.
Closed and open symbols correspond to samples grown on MgO
and Ag, respectively. Dashed lines guide the eye.

Fig. 2. (a) Normalized Raman spectra of amorphous Si layers


measured with excitation at 2.4 eV. The respective average thick-
nesses are given next to the spectra in Å. The spectra were vertically
shifted for clarity. The horizontal solid lines indicate the zero-signal
levels. (b) Raman spectra of the 30 Å sample measured in VV and
VH polarization.

3. Results and discussion

Raman spectra of amorphous samples with a different


average thickness of the Si layer grown on MgO are
shown in Fig. 2(a). As expected, the spectra of the relatively
thick layers are similar to that of amorphous Si (a-Si) and
consist of the TA, LA, and TO strong bands centered around
150, 310 and 480 cm, respectively, and the weak LO
shoulder at 370 cm 21 [29–31], marked in Fig. 2(b). The
Raman spectra are polarized at the linearly polarized excita-
tion. The VV and VH spectra are shown in Fig. 2(b) for the
30 Å sample. The degree of linear polarization, defined as
…IVV 2 IVH †=…IVV 1 IVH †; is about 0.25, which is close to the
value previously determined for a-Si [32]. IVV and IVH are
the Raman intensities in the corresponding configuration.
As the Si layer thickness decreases, the high frequency
half-width-at-half-maximum (HWHM) of the TO band d TO
defined as shown in Fig. 2(a) changes from 35 to 50 cm 21.
At the same time, the position of the maximum D TO moves
Fig. 4. (a) Normalized Raman spectra of the 20 Å sample measured
towards the low Raman shifts. The dependences of D TO and at different temperatures during crystallization process. (b) Spectra
d TO on d are presented in Fig. 3 for samples grown on both of the k ˆ 0 optical phonon in the same sample before (open
MgO and Ag. The observed broadening of the optical band circles) and after termination of the surface with a hydrogen plasma
in the Raman spectra of Si layers corresponds to the modi- (closed circles). Vertical dashed line shows a position of the c-Si
fication of the phonon density of states due to changes in the (100) phonon [35] at 519:7 ^ 0:8 cm21 :
556 A.A. Sirenko et al. / Solid State Communications 113 (2000) 553–558

angular distribution of the sp-bonds in the near-surface plasma, the position of the k ˆ 0 peak shifts back and coin-
atomic layers and dangling bonds on the surface [24,32]. cides within our experimental accuracy with the c-Si (100)
The surface effects are expected to be more pronounced in phonon. The measurements were taken with the same pump-
ultrathin Si layers consisting of amorphous clusters, where a ing power and the heating effect was excluded. Thus, we
fraction of the surface atoms to interior atoms increases conclude that the initially observed red shift of about
significantly. Indeed, the significant variation in the phonon 1.5 cm 21 is determined not by confinement of the optical
spectra occurs for d less than 50 Å, the same thickness-range phonons but is mostly due to the surface strain. In passivated
where the formation of the Si clusters was observed. In the recrystallized samples, hydrogen reduces the strain at the
case of relatively thick layers with d $ 50 A;  any possible nanoparticle surface by destroying unfavorable tensile-
corrugation of the Si surface …Dd ˆ ^10 A†  should not strained Si–Si bonds [36]. A slight decrease in the FWHM
significantly affect the Raman spectra, which are similar of the Raman line (see Fig.4(b)) can be also due to this
to that of the a-Si films. Note, that in the samples grown effect. Further we will briefly discuss effects, which can
on MgO buffer layers, formation of the Si–O bonds at the influence the position of the phonon lines in the Raman
interface could lead to the similar transformation in the spectra of nanocrystals of intermediate size. After that we
Raman spectra, especially in the ultrathin layers, where will proceed to the case of ultrasmall crystalline particles.
the fraction of the surface atoms is high. However, the thick- The absence of the confinement-induced shift of the opti-
ness dependences of the optical band for the samples grown cal phonon in the samples with d between 20 and 50 Å lacks
on both MgO and Ag were identical (see Fig. 3). Thus, we a straightforward interpretation. On one hand, a simple esti-
conclude that the vibration properties of the amorphous Si mation of the confinement-induced frequency shift of the
nanoparticles are determined by the size effects. Our experi- optical phonon in a hemispherical particle [15] results in a
ments are also important to distinguish between size-depen- value of about 5 cm 21 for the sample with d ˆ 20 A;  which
dent modifications of the phonon density of states from that should be measurable with our experimental accuracy. On
due to contamination of the Si surface with hydrogen or the other hand, there are several factors, which can strongly
oxygen. The latter might be neglected for our samples, but affect this simple picture decreasing the expected confine-
could be very strong in, e.g. porous Si. ment-induced effect below our experimental limits:
After heating the samples, a remarkable transformation in
the Raman spectra was observed. Fig. 4(a) shows several 1. The wide size distribution of the nanoparticles, which can
spectra for the 20 Å sample taken at different temperatures. result in a preferable excitation of the quantum dots with
The sharp crystalline peak arises when the temperature the maximum size increasing the average size of the dots,
increases to more than T . 2608C: Note the temperature- which are probed in the Raman experiments.
induced red shift of this Raman line. At the same time, the 2. The inhomogeneous broadening of the Raman line,
intensity of the amorphous-like broad features decreases at which is caused by the higher-order confined modes
T . 3208C: This indicates a partial recovering of the k- with the frequencies lower than the primary confined
vector conservation condition with crystallization of the Si Raman line [37] and by the splitting between TO and
layers allowing only scattering by optical phonons at the LO phonons.
zone center …k ˆ 0†; where the TO and LO phonons in 3. The strain between the dots and the substrate [5,7]. Note
bulk crystalline Si (c-Si) are degenerate [33]. The degree that the compressive strain shifts the line position in the
of linear polarization of this peak is about 0.16, which corre- opposite direction to that induced by confinement.
sponds to the random orientation of the crystallographic 4. An increase of the fraction of the surface to interior atoms
directions in the Si nanoparticles. The crystallization in smaller nanoparticles, leading to modification of the
temperature was found to be around 250–3008C for samples phonon density of states [28,32]. Note that the contribu-
grown on both MgO and Ag (note that the crystallization tion of the surface vibrations may affect the Raman spec-
temperature for a-Si films varies in a wide range depending trum as well [38].
on the sample preparation condition and substrates, see, e.g. The combination of the above-mentioned factors, in prin-
Ref. [34]). ciple, could be responsible for the fact that in the studied Si
The position of the k ˆ 0 peak does not change signifi- clusters, which contain only a few 1000 atoms, no strong
cantly in different recrystallized samples. Its shape is asym- systematic shift of the k ˆ 0 peak was detected. A compar-
metric and a full-width-at-half-maximum (FWHM) varies ison of higher resolution Raman measurements with precise
from 6 cm 21 for thick layers to about 8 cm 21 (see Fig.4(b)) theoretical investigation of the phonon confinement, which
for 20 Å sample. For d # 50 A;  the typical red shift of this should take into account the exact shape of the nanoparticles
line with respect to the position of c-Si (100) phonon (the (see Fig. 1), surface effects, and interaction with substrate,
room temperature value for the k ˆ 0 phonon in c-Si varies are required to clarify the situation. At this point we can only
in literature between 519 and 522 cm 21, see, e.g. Refs. [35]) mention that our data are more similar to results for phonon
is about 1:5 ^ 0:5 cm21 : From the first glance, this shift confinement in thin Si films, where the relatively small
could be attributed to the phonon confinement. However, Raman shifts were observed [15], and to specific cases of
after termination of the sample surface with a hydrogen porous Si [39].
A.A. Sirenko et al. / Solid State Communications 113 (2000) 553–558 557

shown in inset to Fig. 5. In the case of the relatively thick


films, with d greater than 45 Å, the complete k-vector
conservation could be achieved after annealing, which
manifests itself by the vanishing of the amorphous-like
features in the Raman spectra. In contrast, for Si layers
with d less than 10 Å, the Raman spectra do not change
with annealing and practically no modifications have been
observed even after heating up to 6008C for 1 h. Two corre-
sponding spectra of the 10 Å sample grown on a Ag buffer
layer and measured at room temperature before and after
annealing are shown in Fig. 5. This result is in agreement
with the earlier prediction [32] and experimental observa-
tions [23,38,40] of the fact, that in ultrasmall nanocrystals
the k ˆ 0 line vanishes and Raman spectra of both covalent
and ionic nanocrystals become amorphous-like when they
are small enough. Two reasons can be adduced to explain
this effect. First, even in the ideal case of isolated nanocrys-
tal, the decrease in size breaks the crystal periodicity allow-
Fig. 5. Raman spectra of the 10 Å sample before (solid line) and ing observation of the scattering from entire branches of the
after the annealing (dashed line). Inset shows the ratio of the k ˆ 0 acoustic and optical phonons. The situation becomes similar
peak intensity Ikˆ0 to that of the amorphous background Ia as a to a system of weakly interacting molecules in liquids or
function of the average thickness of the Si layers. gases and the consideration of k-vector conservation
becomes needless. Second, the fraction of disordered
Nevertheless, in the ultrathin recrystallized Si layers, we surface atoms increases with respect to the ordered atoms
expected a formation of much smaller Si clusters with a within a crystalline lattice [28,32]. The observed size-
typical size of 10–30 Å, where the effect of the phonon dependent modification of the Raman spectra is important
confinement should become dominant. However, instead for determination of the minimum size of a crystalline
of a larger red shift of the phonon line, the k ˆ 0 peak system, which can satisfy k-vector conservation conditions.
disappeared in samples with d # 10 A:  As the average According to the aforementioned empirical relation between
thickness of the Si layer decreases, the relative intensity of d and the average size of the nanoparticles, the transition
the k ˆ 0 peak to the intensity of the amorphous-like broad between crystalline- and amorphous-like behavior takes
peak at 480 cm 21 becomes smaller. The ratio of the k ˆ 0 place in the recrystallized samples with an average number
peak intensity to the amorphous-like background intensity is of Si atoms in one hemispherical particle equal to …7 ^ 2† ×
102 :
To support this statement an additional verification of the
crystalline structure of the ultrasmall nanoparticles is
required. Note, that one of the alternative explanations of
the observed effect can be an increase of the temperature of
the amorphous-to-crystalline structural transition. However,
our in situ measurements of electron energy losses demon-
strate significant modification of the electronic gap with
annealing in the same temperature range, confirming that
the structure of the annealed Si nanoparticles is crystalline.
The EELS spectra obtained before and after recrystallization
for a d ˆ 5 A  sample grown on SiO2 are shown in Fig. 6.
The intensity was normalized to the elastic peak amplitude.
The onsets of electronic transitions (shown with arrows in
Fig. 6) in the amorphous sample occurs at much lower
energy …0:45 ^ 0:05 eV† than that seen in the annealed
sample …1:55 ^ 0:05 eV; similar values for the optical gap
can be extracted from the EELS spectra by computing the
Fig. 6. EELS spectra of the 5 Å sample before (triangles) and after absorption coefficient and using analogous techniques as
the annealing (circles). Dashed lines guide the eye to the onset of the those employed to amorphous semiconductors, for more
electronic transitions indicated with arrows. Spectra for polycrystal- detail see Ref. [41]). The latter value is much higher than
line Si and SiO2 are shown for comparison with open squares and the indirect gap of bulk crystals [42], which provides a
solid line, respectively. clear evidence of the electron confinement in ultrasmall
558 A.A. Sirenko et al. / Solid State Communications 113 (2000) 553–558

nanocrystals. The observed increase in the electronic gap is [11] T. van Buuren, L.N. Dinh, L.L. Chase, W.J. Siekhaus, L.J.
in a good agreement with the recently reported changes in Terminello, Phys. Rev. Lett. 80 (1998) 3803.
the band ages of silicon nanocrystals by X-ray absorption [12] D. Kovalev, H. Heckler, M. Ben-Chorin, G. Polisski, M.
and photoemission [11]. Schwartzkopff, F. Koch, Phys. Rev. Lett. 81 (1998) 2803.
[13] R. Alben, D. Weaire, J.E. Smith Jr., M.H. Brodsky, Phys. Rev.
B 11 (1975) 2271.
[14] T. Takagahara, J. Lumin. 70 (1996) 129.
4. Conclusions
[15] I.H. Campbell, P.M. Fauchet, Solid State Commun. 58 (1986)
739.
The phonon states of Si nanoparticles free of chemisorbed [16] A. Ekimov, J. Lumin. 70 (1996) 1 and references therein.
species, which were grown and measured in ultrahigh [17] A. Fillos, S.S. Hefner, R. Tsu, J. Vac. Sci. Technol. B 14 (6)
vacuum, have been studied. Remarkable transformation in (1996) 3431.
the phonon density of states with decrease in size of amor- [18] Y. Kanemutsu, H. Uto, Y. Masumoto, T. Masumoto, T.
phous Si nanoparticles was observed. It has been found that Futagi, H. Mimura, Phys. Rev. B 48 (1993) 2827.
the surface strain induced by the dangling bonds can result [19] S. Guha, P. Steiner, W. Lang, J. Appl. Phys. 79 (1996) 8664.
in a red shift of the optical phonon frequency for relatively [20] Z. Igbal, S. Veprek, J. Phys. C: Solid State Phys. 15 (1982)
big nanocrystals. We consider the observed relaxation of the 377.
k-vector conservation in ultrasmall nanocrystals as the most [21] H. Tanino, A. Kuprin, H. Deai, N. Koshida, Phys. Rev. B 53
(1996) 1937.
important result of this paper. The average number of Si
[22] Y. Gao, T. López-Rı́os, Solid State Commun. 60 (1986) 55.
atoms per one particle, which corresponds to the transition
[23] S. Hayashi, H. Abe, Jpn. J. Appl. Phys. 23 (1984) L824.
between crystalline- and amorphous-like behavior, has been [24] J. Fortner, J.S. Lannin, Surf. Sci. 254 (1991) 251.
determined. A confinement-induced increase in the electro- [25] G.A.N. Connel, R.J. Nemanich, C.C. Tsai, Appl. Phys. Lett.
nic gap of ultrasmall Si particles has been measured by 36 (1980) 31.
means of in situ EELS providing a clear evidence of their [26] W.S. Bacsa, J.S. Lannin, Appl. Phys. Lett. 61 (1992) 19.
crystalline structure. [27] J.C. Tsang, in: M. Cardona, G. Guntherodt (Eds.), Light Scat-
tering in Solids, Topics in Applied Physics, Vol. V, Springer,
Berlin, 1989, p. 233, chap. 6.
Acknowledgements [28] V.I. Merkulov, J.S. Lannin, J.M. Cowley, in: R.W. Collins
(Ed.), Advances in Microcrystalline and Nanocrystalline
Semiconductors, Materials Research Society Symposia
The authors are grateful to V.I. Merkulov for useful
Proceedings No. 452MRS, Pittsburgh, 1996, p. 225.
discussions. This work was supported by USDOE Grant [29] J.E. Smith Jr., M.H. Brodsky, B.L. Crowder, M.I. Nathan, A.
DE-FG02-84ER45095 and NSF Grant DMR-96-23315. Pinczuk, Phys. Rev. Lett. 26 (1971) 642.
[30] J.S. Lannin, P.J. Carrol, Philos. Mag. 45 (1982) 155.
[31] G. Nilson, G. Nelin, Phys. Rev. B 6 (1972) 3777.
References [32] J.S. Lannin, in: J.I. Pankove (Ed.), Semiconductors and Semi-
metals, 21(B), Academic Press, Orlando, 1984, p. 159.
[1] A.G. Cullis, L.T. Canham, Nature (London) 353 (1991) 353. [33] P.Y. Yu, M. Cardona, Fundamentals of Semiconductors,
[2] A.G. Cullis, L.T. Canham, P.D.J. Calcott, J. Appl. Phys. 82 Springer, Berlin, 1996, p. 103.
(1997) 909 and references therein. [34] D. Bimberg, et al., in: O. Madelung, et al. (Eds.), Numerical
[3] R. Tsu, J. Morais, Bowhill, in: Z. Feng, R. Tsu (Eds.), Porous data and functional relationships in science and technology,
Silicon, World Scientific, Singapore, 1994, p. 443. Landolt-Bornstein, New Series, Group III, Vol. 17i, Springer,
[4] K. Eberl, K. Brunner, W. Winter, Thin Solid Films 294 (1997) Berlin, 1985, p. 5.
98. [35] D. Bimberg, et al., in: O. Madelung, et al. (Eds.), Numerical
[5] Z. Igbal, S. Veprek, A.P. Webb, P. Capezzuto, Solid State Data and Functional Relationships in Science And Technol-
Commun. 37 (1981) 993. ogy, Landolt-Bornstein, New Series, Group III, Vol. 17a,
[6] E.W. Forsythe, E.A. Whittaker, F.H. Pollak, S.W. Sywe, G.S. Springer, Berlin, 1982, p. 72.
Tompa, B.A. Khan, J. Khurgin, H.W.H. Lee, F. Adar, H. [36] I. Kaiser, N.H. Nickel, W. Fuhs, W. Pilz, Phys. Rev. B 58
Schaffer, in: R.W. Collins (Ed.), Microcrystalline and Nano- (1998) R1718.
crystalline Semiconductors, Materials Research Society [37] C. Trallero-Giner, A. Debernardi, M. Cardona, A.I. Ekimov,
Symposia Proceedings No. 358MRS, Pittsburgh, 1994, p. 187. Phys. Rev. B 57 (1998) 4664.
[7] X.S. Zhao, Y.R. Ge, J. Schroeder, P.D. Persants, Appl. Phys. [38] T. Okada, T. Iwaki, K. Yamamoto, H. Kashara, K. Abe, Solid
Lett. 65 (1994) 2033. State Commun. 49 (1984) 809.
[8] L. Tsybeskov, MRS Bull. 23 (4) (1998) 33 and references [39] F. Kozlowski, P. Steiner, F. Lang, in: Z. Feng, R. Tsu (Eds.),
therein. Porous Silicon, World Scientific, Singapore, 1994, p. 149.
[9] W.L. Wilson, P.F. Szajowski, L.E. Brus, Science 262 (1993) [40] S. Hayashi, Solid State Commun. 56 (1985) 375.
1242. [41] G.P. Lopinski, J.S. Lannin, Appl. Phys. Lett. 69 (1996)
[10] S. Schlupper, S.L. Friedman, M.A. Marcus, D.L. Adler, Y.-H. 2400.
Xie, F.M. Ross, T.D. Harris, W.L. Brown, Y.J. Chabal, L.E. [42] G.D. Cody, in: J.I. Pankove (Ed.), Semiconductors and Semi-
Brus, P.H. Citrin, Phys. Rev. Lett. 72 (1994) 2648. metals, Vol. 21B, Academic Press, Orlando, 1984, p. 11.

Das könnte Ihnen auch gefallen